1 Introduction

Multipliers of algebras, particularly multipliers of Banach algebras, have been studied in the field of analysis. In this paper we discuss them in a purely algebraic manner.

Let A be a Banach algebra. A mapping \(T: A \rightarrow A\) is termed a multiplier of A if it satisfies the condition (I) \(xT(y) = T(xy) = T(x)y\) for all \(x, y \in A\). We denote the collection of all multipliers of A as M(A), and the collection of all bounded linear operators on A as B(A). Notably, M(A) forms an algebra and B(A) constitutes a Banach algebra. A Banach algebra A is referred to as without order if it has neither a nonzero left annihilator nor a nonzero right annihilator. If A is without order, the algebra M(A) forms a commutative closed subalgebra of B(A) (see [2], Proposition 1.4.11). In 1952, Wendel [8] proved an important result that the multiplier algebra of \(L^1(G)\) on a locally compact group G is isometrically isomorphic to the measure algebra on G. The general theory of multipliers of Banach algebras has been developed by Johnson [1]. For a comprehensive reference on the theory of multipliers of Banach algebras, refer to Larsen [5].

When A is without order, T is a multiplier if it satisfies the condition (II) \(xT(y) = T(x)y\) for all \(x, y \in A\). Many researchers had been unaware of the difference between conditions (I) and (II) until Zivari-Kazempour [9] (see also [10]) recently articulated the difference. We call a mapping T satisfying (II) a weak multiplier and denote the set of such multipliers of A by \(M'(A)\). Then, M(A) is in general a proper subset of \(M'(A)\). Furthermore, (weak) multipliers can be defined for an algebra A that is not necessarily associative, and they are not linear mappings in general. We denote the spaces of linear multipliers and linear weak multipliers of A by LM(A) and \(LM'(A)\) respectively. M(A) and LM(A) are subalgebras of the algebra \(A^A\) consisting of all mappings from A to itself. Meanwhile, \(M'(A)\) and \(LM'(A)\) are closed under the operation \(\circ \) defined by \(T \circ S = TS + ST\), and they form a Jordan algebra.

In Sects. 2 to 5, we study general properties of (weak) multipliers. More specifically, in Sects. 3 and 4 we give a decomposition theorem (Theorem 3.1), and a matrix equation (Theorem 4.2) for (weak) multipliers. They play an essential role in our examination of (weak) multipliers.

Complete classifications of associative algebras and zeropotent algebras of dimension 3 over an algebraically closed field of characteristic not equal to 2 were given in Kobayashi et al., [3] and [4]. In Sects. 6 and 7 we undertake a complete determination of the (linear) (weak) multipliers of these algebras.

Some authors have considered other weaker concepts related to multipliers, such as (pseudo-)n-multipliers (for more information, see [6] and [11]).

2 Multipliers and weak multipliers

Let K be a field and A be a (not necessarily associative) algebra over K. The set \(A^A\) of all mappings from A to A forms an associative algebra over K in the usual way. Let L(A) denote the subalgebra of \(A^A\) of all linear mappings from A to A.

A mapping \(T: A \rightarrow A\) is a weak multiplier of A, if

$$\begin{aligned} xT(y) = T(x)y \end{aligned}$$
(1)

holds for any \(x, y \in A\), and T is a multiplier, if

$$\begin{aligned} xT(y) = T(xy) = T(x)y \end{aligned}$$
(2)

for any \(x, y \in A\). Let M(A) (resp. \(M'(A)\)) denote the set of all multipliers (resp. weak multipliers) of A. Define

$$\begin{aligned} LM(A) \overset{\text {\tiny def}}{=}M(A) \cap L(A)\ \, \text{ and }\ \, LM'(A) \overset{\text {\tiny def}}{=}M'(A) \cap L(A). \end{aligned}$$

Proposition 2.1

M(A) (resp. LM(A)) is a unital subalgebra of \(A^A\) (resp. L(A)), and \(M'(A)\) (resp. \(LM'(A)\)) is a Jordan subalgebra of \(A^A\) (resp. L(A)).

Proof

First, the zero mapping is a multiplier because all three terms in (2) are zero. Secondly, the identity mapping is also a multiplier because the three terms in (2) are xy. Let \(T, U \in M(A)\). Then we have

$$\begin{aligned} x(T+U)(y)= & {} xT(y) + xU(y) = T(xy) + U(xy) = T(x)y + U(x)y\nonumber \\= & {} (T+U)(x)y \end{aligned}$$
(3)

and

$$\begin{aligned} x(TU)(y) = xT(U(y)) = T(xU(y)) = TU(xy) = T( U(x)y) = (TU)(x)y\nonumber \\ \end{aligned}$$
(4)

for any \(x, y \in A\). Hence, \(T+U, TU\) belong to M(A). Finally let \(k \in K\), then

$$\begin{aligned} x(kT)(y) = kxT(y) = kT(xy) = kT(x)y = (kT)(x)y, \end{aligned}$$
(5)

and so \(kT \in M(A)\). Therefore, M(A) is a unital subalgebra of \(A^A\), and \(LM(A) = M(A)\cap L(A)\) is a unital subalgebra of L(A).

Next, let \(T, U \in M'(A)\). Then, the equalities in (3) and (5) hold, removing the center terms \(T(xy)+U(xy)\) and kT(xy), respectively. Hence, \(M'(A)\) is a subspace of \(A^A\). Moreover, we have

$$\begin{aligned} x(TU)(y) = xT(U(y)) = T(x)U(y) = U(T(x))y =UT(x)y \end{aligned}$$

and similarly \(x(UT)(y) = TU(x)y\) for any \(x, y \in A\). Hence,

$$\begin{aligned} x(TU + UT)(y) = (TU + UT)(x)y. \end{aligned}$$

It follows that \(TU + UT \in M'(A)\).Footnote 1\(\square \)

Let Ann\(_l(A)\) (resp. Ann\(_r(A)\)) be the left (resp. right) annihilator of A and let \(A_0\) be their intersection, that is,

$$\begin{aligned} \textrm{Ann}_l(A) = \{a \in A\,|\, ax = 0 {\ \mathrm for\ all\ } x \in A\}, \\ \textrm{Ann}_r(A) = \{a \in A\,|\, xa = 0 {\ \mathrm for\ all\ } x \in A\} \end{aligned}$$

and

$$\begin{aligned} A_0 = \textrm{Ann}_l(A) \cap \textrm{Ann}_r(A). \end{aligned}$$

They are all subspaces of A, and when A is an associative algebra, they are two-sided ideals. For a subset X of A, \(\langle X\rangle \) denotes the subspace of A generated by X.

Proposition 2.2

A weak multiplier T of A such that \(\langle T(A)\rangle \cap A_0 = \{0\}\) is a linear mapping.

Proof

Let \(x, y, z \in A\) and \(a, b \in K\), and let T be a weak multiplier. We have

$$\begin{aligned} T(ax + by)z&= (ax+by)T(z) = axT(z) + byT(z) = aT(x)z + bT(y)z\\&= \left( aT(x) + bT(y)\right) z. \end{aligned}$$

Because z is arbitrary, we have \(w = T(ax+by) - aT(x) - bT(y) \in \textrm{Ann}_l(A)\). Similarly, we can show \(w \in \textrm{Ann}_r(A)\), and so \(w \in A_0\). Hence, if \(\langle T(A)\rangle \cap A_0 = \{0\}\), then \(w = 0\) because \(w \in \langle T(A)\rangle \). Since abxy are arbitrary, T is a linear mapping. \(\square \)

Corollary 2.3

If \(A_0 = \{0\}\), then any weak multiplier is a linear mapping over K, that is, \(M'(A) = LM'(A)\) and \(M(A) = LM(A)\).

Proposition 2.4

If T is a weak multiplier, then \(T\left( \textrm{Ann}_l(A)\right) \subseteq \textrm{Ann}_l(A)\), \(T\left( \textrm{Ann}_r(A)\right) \subseteq \textrm{Ann}_r(A)\) and \(T(A_0) \subseteq A_0\).

Proof

Let \(x \in \textrm{Ann}_l(A)\), then for any \(y \in A\) we have

$$\begin{aligned} 0 = xT(y) = T(x)y. \end{aligned}$$

Hence, \(T(x) \in \textrm{Ann}_l(A)\). The other cases are similar. \(\square \)

In this paper we denote the subset \(\{xy\,|\,x, y \in A\}\) of A by \(A^2\).Footnote 2

Proposition 2.5

Any mapping \(T\hspace{-.6mm}: A \rightarrow A\) such that \(T(A) \subseteq A_0\) is a weak multiplier. Such a mapping T is a multiplier if and only if \(T(A^2) = \{0\}\). In particular, if A is the zero algebra, every mapping T is a weak multiplier, and it is a multiplier if and only if \(T(0) = 0\).

Proof

If \(T(A) \subseteq A_0\), then both sides are 0 in (1) and T is a weak multiplier. This T is a multiplier if and only if the term T(xy) in the middle of (2) is 0 for all \(x, y \in A\), that is, \(T(A^2) = \{0\}\). If A is the zero algebra, then \(A = A_0\) and \(A^2\) = {0}. Hence, any T is a weak multiplier and it is a multiplier if and only if \(T(0) = 0\). \(\square \)

The opposite \(A^\textrm{op}\) of A is the algebra with the same elements and addition as A, but the multiplication \(*\) in it is reversed, that is, \(x*y = yx\) for all \(x, y \in A\).

Proposition 2.6

A and \(A^\textrm{op}\) have the same multipliers and weak multiplies, that is,

$$\begin{aligned} M(A^\textrm{op}) = M(A)\ \ \text{ and }\ \ M'(A^\textrm{op}) = M'(A). \end{aligned}$$

Proof

Let \(T \in A^A\). Then, \(T \in M'(A)\) if and only if

$$\begin{aligned} x*T(y) = T(y)x = yT(x) = T(x)*y \end{aligned}$$

for any \(x, y \in A\), if and only if \(T \in M'(A^\textrm{op})\). Further, \(T \in M(A)\) if and only if

$$\begin{aligned} x*T(y) = T(y)x = T(yx) = T(x*y) = yT(x) = T(x)*y \end{aligned}$$

for any \(x, y \in A\), if and only if \(T \in M(A^\textrm{op})\). \(\square \)

3 Nihil decomposition

Let \(A_1\) be a subspace of A such that

$$\begin{aligned} A = A_1 \oplus A_0. \end{aligned}$$
(6)

Here, \(A_1\) is not unique, but choosing an appropriate \(A_1\) will become important later. When \(A_1\) is fixed, any mapping \(T \in A^A\) is uniquely decomposed as

$$\begin{aligned} T = T_1 + T_0 \end{aligned}$$
(7)

with \(T_1(A) \subseteq A_1\) and \(T_0(A) \subseteq A_0\). We call (6) and (7) a nihil decomposition of A and T, respectively. Let \(\pi : A \rightarrow A_1\) be the projection and \(\mu : A_1 \rightarrow A\) be the embedding, that is, \(\pi (x_1+x_0) = \mu (x_1) = x_1\) for \(x_1 \in A_1\) and \(x_0 \in A_0\).

Let \(M_1(A)\) (resp. \(M_0(A)\)) denote the set of all multipliers T of A with \(T(A) \subseteq A_1\) (resp. \(T(A) \subseteq A_0\)). Similarly, the sets \(M'_1(A)\) and \(M'_0(A)\) of weak multipliers of A are defined. Also, set

$$\begin{aligned} LM_i(A) = M_i(A) \cap L(A)\ \,\text{ and }\ \,LM'_i(A) = M'_i(A) \cap L(A) \end{aligned}$$

for \(i = 0,1\). By Proposition 2.2 we see

$$\begin{aligned} M'_1(A) = LM'_1(A)\ \,\text{ and }\ \,M_1(A) = LM_1(A), \end{aligned}$$

and by Proposition 2.5 we have

$$\begin{aligned} M'_0(A) = A_0^A\,\ \text{ and }\,\ M_0(A) = \{T \in A_0^A\,|\,T(A^2) = \{0\}\}. \end{aligned}$$
(8)

Theorem 3.1

Let A = \(A_1 \oplus A_0\) and \(T = T_1 + T_0\) be nihil decompositions of A and \(T \in A^A\) respectively.

(i) T is a weak multiplier, if and only if \(T_1\) is a weak multiplier. If T is a weak multiplier, \(T_1\) is a linear mapping satisfying \(T_1(A_0) = \{0\}\).

(ii) If \(T_1\) is a multiplier and \(T_0(A^2) = \{0\}\), then T is a multiplier. If \(A_1\) is a subalgebra of A, the converse is also true.

Suppose that \(A_1\) is a subalgebra of A, and let \(\Phi \) be a mapping sending \(R \in (A_1)^{A_1}\) to \(\mu \circ R \circ \pi \in A^A\). Then,

(iii) \(\Phi \) gives an algebra isomorphism from \(M(A_1)\) onto \(M_1(A)\) and a Jordan isomorphism from \(M'(A_1)\) onto \(M'_1(A)\).

Proof

Let \(x, y \in A\).

(i) If T is a weak multiplier, then

$$\begin{aligned} xT_1(y) = x(T(y) - T_0(y)) = xT(y) = T(x)y = T_1(x)y. \end{aligned}$$

Thus, \(T_1\) is also a weak multiplier. Moreover, \(T_1\) is a linear mapping by Proposition 2.2 and \(T_1(A_0) \subseteq A_1 \cap A_0 = \{0\}\) by Proposition 2.4. Conversely, if \(T_1\) is a weak multiplier, then

$$\begin{aligned} xT(y) = xT_1(y) = T_1(x)y = T(x)y, \end{aligned}$$

and so T is a weak multiplier.

(ii) If \(T_1\) is a multiplier and \(T_0(A^2) = 0\), then T is a multiplier because

$$\begin{aligned} xT(y) = xT_1(y) = T_1(xy) = T(xy) - T_0(xy) = T(xy) = T(x)y. \end{aligned}$$

Next suppose that \(A_1\) is a subalgebra. If T is a multiplier, then for any \(x, y \in A\) we have

$$\begin{aligned} T_1(xy) + T_0(xy) = T(xy) = xT(y) = x_1T_1(y), \end{aligned}$$
(9)

where \(x = x_1+x_0\) with \(x_1 \in A_1\) and \(x_0 \in A_0\). Here, \(x_1T_1(y) \in A_1\) because \(A_1\) is a subalgebra, and thus, we have \(T_0(xy) = x_1T_1(y) - T_1(xy) \in A_0 \cap A_1 = \{0\}\). Since xy are arbitrary, we get \(T_0(A^2) = \{0\}\). Moreover, because \(T_1(xy) = x_1T_1(y) = xT_1(y)\) by (9) and similarly \(T_1(xy) = T_1(x)y\), \(T_1\) is a multiplier. The converse is already proved above.

(iii) Let \(S \in (A_1)^{A_1}\) and \(x = x_1+x_0, y = y_1 + y_0 \in A\) with \(x_1, y_1 \in A_1\) and \(x_0, y_0 \in A_0\). Then, \(\pi (x) = \mu (x_1) = x_1\), \(\pi (y) = \mu (y_1) = y_1\) and

$$\begin{aligned} \Phi (S)(x) = \mu (S(\pi (x))) = \mu (S(x_1)) = S(x_1). \end{aligned}$$

Thus, if \(S \in M'(A_1)\), we have

$$\begin{aligned} x\Phi (S)(y) = xS(y_1) = x_1S(y_1) = S(x_1)y_1 = \Phi (S)(x)y_1 = \Phi (S)(x)y. \end{aligned}$$

Hence, \(\Phi (S) \in M'_1(A)\). Moreover, if \(S \in M(A_1)\), then because \(\pi \) is a homomorphism, we have

$$\begin{aligned} \Phi (S)(xy) = S(\pi (xy)) = S(x_1y_1) = x_1S(y_1) = x\Phi (S)(y), \end{aligned}$$

and so \(\Phi (S) \in M_1(A)\).

Conversely, let \(T \in M'_1(A)\), then because T is a linear mapping satisfying \(T(A_0) = \{0\}\), there is a linear mapping \(S \in L(A_1)\) on \(A_1\) such that \(\Phi (S) = T\), that is, \(S(x_1) = T(x) = T(x_1)\). We have

$$\begin{aligned} x_1S(y_1) = x_1T(y_1) = T(x_1)y_1 = S(x_1)y_1, \end{aligned}$$

and hence \(S \in M'(A_1)\). Therefore, \(\Phi \) induces a linear isomorphism from \(M'(A_1)\) to \(M'_1(A)\). Similarly, \(\Phi \) gives a linear isomorphism from \(M(A_1)\) to \(M_1(A)\). Moreover, for \(T, U \in M'(A_1)\), we have

$$\begin{aligned} \Phi (TU) = \mu \circ T\circ U\circ \pi = \mu \circ T\circ \pi \circ \mu \circ U\circ \pi = \Phi (T)\Phi (U). \end{aligned}$$

Thus, \(\Phi \) gives an isomorphism of algebras from \(M(A_1)\) to \(M_1(A)\) and a Jordan isomorphism from \(M'(A_1)\) to \(M'_1(A)\). \(\square \)

Theorem 3.1 implies

$$\begin{aligned} M'(A) = \ M'_1(A) \oplus M'_0(A)\ \text{ and }\ M_1(A) \oplus M_0(A) \subseteq M(A), \end{aligned}$$

where \(M'_0(A)\) and \(M_0(A)\) are given as (8). If \(A_1\) is a subalgebra, we have

$$\begin{aligned} M'(A) \cong M'(A_1) \oplus (A_0)^A\ \text{ and }\ M(A) \cong M(A_1) \oplus \left\{ T \in (A_0)^A\,|\,T(A^2) = \{0\}\right\} .\nonumber \\ \end{aligned}$$
(10)

Corollary 3.2

Any weak multiplier T is written as

$$\begin{aligned} T = T_1 + R \end{aligned}$$
(11)

with \(T_1 \in LM_1'(A)\) and \(R \in (A_0)^{A},\) and it is a multiplier if and only if

$$\begin{aligned} R(x_1y_1) = x_1T_1(y_1) - T_1(x_1y_1) \end{aligned}$$
(12)

for any \(x_1, y_1 \in A_1\).

Proof

As stated above T is written as (11). Let \(x = x_1+x_0, y = y_1 + y_0 \in A\) with \(x_1, y_1 \in A_1\) and \(x_0, y_0 \in A_0\) be arbitrary, then we have

$$\begin{aligned} xT(y) = x_1(T_1(y) + R(y)) = x_1T_1(y) = x_1T_1(y_1) \end{aligned}$$
(13)

because \(R(A) \subseteq A_0\) and \(T_1(A_0) = \{0\}\). The last term in (13) is also equal to \(T_1(x_1)y_1 = T(x)y\). Hence, T is a multiplier if and only if it is equal to \(T(xy) = T(x_1y_1) = T_1(x_1y_1) + R(x_1y_1)\), if and only if (12) holds. \(\square \)

4 Linear multipliers and matrix equation

In this section, A is a finite-dimensional algebra over K. We derive a matrix equation that characterizes a (weak) multiplier for a linear mapping on A. Suppose that A is n-dimensional with basis \(E = \{e_1, e_2, \dots , e_n\}\).

Lemma 4.1

A linear mapping \(T: A \rightarrow A\) is a weak multiplier if and only if

$$\begin{aligned} e_iT(e_j) = T(e_i)e_j, \end{aligned}$$
(14)

and it is a multiplier if and only if

$$\begin{aligned} T(e_ie_j) = e_iT(e_j) = T(e_i)e_j, \end{aligned}$$
(15)

for all \(e_i, e_j \in E\).

Proof

The necessity of the conditions (14) and (15) is obvious. Let \(x = x_1e_1 + x_2e_2+ \cdots + x_ne_n, y = y_1e_1+y_2e_2+\cdots +y_ne_n \in A\) with \(x_1,x_2,\dots ,x_n, y_1,y_2,\dots ,y_n \in K\). If T satisfies (14), then we have

$$\begin{aligned} xT(y)= & {} \left( \sum _i x_ie_i\right) T\left( \sum _j y_je_j\right) = \left( \sum _i x_ie_i\right) \left( \sum _j y_jT(e_j)\right) \\= & {} \sum _{i,j} x_iy_je_iT(e_j) =\sum _{i,j}x_iy_jT(e_i)e_j\\= & {} \left( \sum _ix_iT(e_i)\right) \left( \sum _jy_je_j\right) = T(x)y. \end{aligned}$$

Hence, T is a weak multiplier. Moreover, if T satisfies (15), it is a multiplier in a similar manner. \(\square \)

Let \(\varvec{A}\) (we use the bold character) represent the multiplication table of A on E. \(\varvec{A}\) is a matrix whose elements are drawn from A and given by

$$\begin{aligned} \varvec{A} = \varvec{E}^t\varvec{E}, \end{aligned}$$
(16)

where \(\varvec{E} = (e_1, e_2, \dots , e_n)\) (we again use the boldface \(\varvec{E}\)) is the row vector consisting the basis elements and \(\varvec{E}^t\) is its transpose. For a linear mapping T on A and a matrix \(\varvec{B}\) over A, \(T(\varvec{B})\) denotes the matrix obtained by applying T element-wise, that is, the (ij)-element of \(T(\varvec{B})\) is \(T(b_{ij})\) for the (ij)-element \(b_{ij}\) of \(\varvec{B}\).Footnote 3 We employ the same symbol T for the representation matrix of T on E, that is,

$$\begin{aligned} T(\varvec{E}) = \varvec{E}T. \end{aligned}$$
(17)

Theorem 4.2

A linear mapping T is a weak multiplier of A if and only if

$$\begin{aligned} \varvec{A}T = T^t\varvec{A}, \end{aligned}$$
(18)

and T is a multiplier if and only if

$$\begin{aligned} T(\varvec{A}) = \varvec{A}T = T^t\varvec{A}. \end{aligned}$$
(19)

Proof

By (16) and (17) we have

$$\begin{aligned} (e_1, e_2, \dots , e_n)^t\left( T(e_1), T(e_2), \dots , T(e_n)\right) = \varvec{E}^tT(\varvec{E}) = \varvec{E}^t\varvec{E}T = \varvec{A}T \end{aligned}$$
(20)

and

$$\begin{aligned} \left( T(e_1), T(e_2), \dots , T(e_2)\right) ^t(e_1, e_2, \dots , e_n) = T(\varvec{E})^t\varvec{E} = T^t\varvec{E}^t\varvec{E} = T^t\varvec{A}. \end{aligned}$$
(21)

By Lemma 4.1, T is a weak multiplier if and only if (20) and (21) are equal, if and only if (18) holds. Moreover, T is multiplier if and only if the leftmost sides of (20) and (21) are equal to \((T(e_ie_j))_{i,j=1,2,\dots ,n} = T(\varvec{A})\), if and only if (19) holds. \(\square \)

The multiplication table of the opposite algebra \(A^\textrm{op}\) of A is the transpose \(\varvec{A}^\textrm{t}\) of \(\varvec{A}\). So, the algebras with multiplication tables transposed to each other share the same (weak) multipliers.

5 Associative algebras

In this section, A is an associative algebra over K.

Proposition 5.1

If \(A_0 = \{0\}\), then we have

$$\begin{aligned} M(A) = M'(A) = LM(A) = LM'(A). \end{aligned}$$

Proof

Let \(T \in M'(A)\), then we have

$$\begin{aligned} T(xy)z = xyT(z) = xT(y)z\ \ \textrm{and}\ \ zT(xy) = T(z)xy = zT(x)y \end{aligned}$$

for any \(x, y, z \in A\). It follows that

$$\begin{aligned} T(xy) - xT(y) \in \textrm{Ann}_l(A) \cap \textrm{Ann}_r(A) = A_0 =\{0\}. \end{aligned}$$

Hence, \(T(xy) = xT(y)\) and we see \(T \in M(A)\). Moreover, \(T \in LM(A)\) by Proposition 2.2. \(\square \)

Let \(a \in A\). If \(xay = axy\) (resp. \(xay = xya\)) for any \(x, y \in A\), a is called a left (resp. right) central element, and we simply call it a central element if \(ax = xa\) for any \(x \in A\). Let \(Z_l(A)\), (resp. \(Z_r(A)\), Z(A)) denote the set of all left central (resp. right central, central) elements.

For \(a\in A\), \(l_a\) (resp. \(r_a\)) denotes the left (resp. right) multiplication by a, that is,

$$\begin{aligned} l_a(x) = ax,\ \ \ r_a(x) = xa \end{aligned}$$

for \(x \in A\). They are linear mappings.

Lemma 5.2

For \(a \in A\) the following statements are equivalent.

(i) \(l_a\) (resp. \(r_a\)) is a multiplier,

(ii) \(l_a\) (resp. \(r_a\)) is a weak multiplier,

(iii) a is left (resp. right) central.

Proof

If \(l_a\) is a weak multiplier, then

$$\begin{aligned} xay = xl_a(y) = l_a(x)y = axy \end{aligned}$$

for any \(x, y \in A\), which implies that a is left central. Conversely, if a is left central, \(l_a\) is a weak multiplier also by the above equalities. Moreover, \(l_a\) is a multiplier because \(l_a(xy) = axy = l_a(x)y\). The other case is analogous, and we see that these three statements are equivalent. \(\square \)

As can be easily proved, \(Z_l(A)\) (resp. \(Z_r(A)\)) is a subalgebra of A containing \(\textrm{Ann}_l(A)\) (resp. \(\textrm{Ann}_r(A)\)). Hence, we can form the quotient algebras \(\bar{Z}_l(A) = Z_l(A)/ \textrm{Ann}_l(A)\) and \(\bar{Z}_r(A) = Z_r(A)/\textrm{Ann}_r(A)\).

Theorem 5.3

Suppose that A has a left (resp. right) identity e. Then, any weak multiplier is a left (resp. right) multiplication by a left (resp. right) central element and it is a linear multiplier. The mapping \(\phi \hspace{-.6mm}: Z_l(A)\ (resp.\ Z_r(A)) \rightarrow M'(A)= LM(A)\) sending \(a \in Z_l(A)\ (resp.\ Z_r(A))\) to \(l_a\) (resp. \(l_r\)) induces an isomorphism \(\bar{\phi }\hspace{-.6mm}: \overline{Z_l}(A)\ (resp.\ \overline{Z_r}(A)) \rightarrow M(A)\) of algebras. In particular, if A is unital, M(A) is isomorphic to Z(A).

Proof

Suppose that A has a left identity e. Let \(T \in M'(A)\) and set \(a = T(e)\). Then we have

$$\begin{aligned} T(x) = eT(x) = T(e)x = ax \end{aligned}$$

for any \(x \in A\). Hence, \(T = l_a\), where \(a \in Z_l(A)\) and T is a linear multiplier by Lemma 5.2. Therefore, \(M'(A) = LM(A)\) and \(\phi \) is surjective. Moreover, for \(a \in Z_l(A)\), \(\phi (a) = 0\) if and only if \(ax = 0\) for any \(x \in A\), if and only if \(a \in \textrm{Ann}_l(A)\). Thus we have Ker\((\phi ) = \textrm{Ann}_l(A)\), and \(\phi \) induces the desired isomorphism. Similarly, if A has a right identity, M(A) is isomorphic to \(\overline{Z_r}(A)\). Lastly, if A has the identity, then \(Z_\ell (A) = Z(A)\) and Ann\(_l(A) = \{0\}\), and hence M(A) is isomorphic to Z(A). \(\square \)

6 3-dimensional associative algebras

Over an algebraically closed field K of characteristic not equal to 2, we have, up to isomorphism, 24 families of associative algebras of dimension 3. They consist of 5 unital algebras \(U_0, U_1, U_2\), \(U_3, U_4\) defined on basis \(E = \{e,f,g\}\) by the multiplication tables

$$\begin{aligned} \begin{pmatrix} e&{} \quad f&{} \quad g\\ f&{} \quad 0&{} \quad 0\\ g&{} \quad 0&{} \quad 0 \end{pmatrix}, \ \begin{pmatrix} e&{} \quad f&{} \quad g\\ f&{} \quad 0&{} \quad f\\ g&{} \quad -f&{} \quad e \end{pmatrix}, \ \begin{pmatrix} e&{} \quad 0&{} \quad 0\\ 0&{} \quad f&{} \quad 0\\ 0&{} \quad 0&{} \quad g \end{pmatrix}, \ \begin{pmatrix} e&{} \quad 0&{} \quad 0\\ 0&{} \quad f&{} \quad g\\ 0&{} \quad g&{} \quad 0 \end{pmatrix}, \ \begin{pmatrix} e&{} \quad f&{} \quad g\\ f&{} \quad g&{} \quad 0\\ g&{} \quad 0&{} \quad 0 \end{pmatrix}, \end{aligned}$$

5 curled algebras \(C_0, C_1, C_2, C_3, C_4\) defined by

$$\begin{aligned}{ \begin{pmatrix} 0 &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 0 \end{pmatrix}, \ \begin{pmatrix} 0 &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad e \\ 0 &{} \quad -e &{} \quad 0 \end{pmatrix}, \ \begin{pmatrix} {0} &{} \quad {0} &{} \quad {0} \\ {e} &{} \quad {f} &{} \quad {0} \\ {0} &{} \quad {g} &{} \quad {0} \end{pmatrix}, \ \begin{pmatrix} {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {0} \\ {e} &{} \quad {f} &{} \quad {g} \end{pmatrix}, \ \begin{pmatrix} {0} &{} \quad {0} &{} \quad {e} \\ {0} &{} \quad {0} &{} \quad {f} \\ {0} &{} \quad {0} &{} \quad {g} \end{pmatrix},} \end{aligned}$$

non-unital 4 straight algebras \(S_1, S_2, S_3, S_4\) defined by

$$\begin{aligned} {\begin{pmatrix} {f}\ {} &{} \quad {g} &{} \quad {0} \\ {g}\ {} &{} \quad {0} &{} \quad {0} \\ {0}\ {} &{} \quad {0} &{} \quad {0} \end{pmatrix},\ \begin{pmatrix} {e}\ {} &{} \quad {0} &{} \quad {0} \\ {0}\ {} &{} \quad {g} &{} \quad {0} \\ {0}\ {} &{} \quad {0} &{} \quad {0} \end{pmatrix},\ \begin{pmatrix} {e}\ {} &{} \quad {0} &{} \quad {0} \\ {0}\ {} &{} \quad {f} &{} \quad {0} \\ {0}\ {} &{} \quad {0} &{} \quad {0} \end{pmatrix},\ \begin{pmatrix} {e}\ {} &{} \quad {f} &{} \quad {0} \\ {f}\ {} &{} \quad {0} &{} \quad {0} \\ {0}\ {} &{} \quad {0} &{} \quad {0} \end{pmatrix},} \end{aligned}$$

and non-unital 10 families of waved algebras \(W_1\), \(W_2\), \(W_4\), \(W_5\), \(W_6\), \(W_7\), \(W_8\), \(W_9\), \(W_{10}\) and \(\big \{W_3(k)\big \}_{k \in H}\) defined by

$$\begin{aligned} {\begin{pmatrix} {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {e} \end{pmatrix},\ \begin{pmatrix} {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {e} &{} \quad {0} \end{pmatrix},\ \begin{pmatrix} {e} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {0} \end{pmatrix},\ \begin{pmatrix} {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {f} &{} \quad {g} \end{pmatrix},\ \begin{pmatrix} {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {f} \\ {0} &{} \quad {0} &{} \quad {g} \end{pmatrix},} \end{aligned}$$
$$\begin{aligned} { \begin{pmatrix} e &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad f &{} \quad g \end{pmatrix},\ \begin{pmatrix} {e} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {0} &{} \quad {f} \\ {0} &{} \quad {0} &{} \quad {g} \end{pmatrix},\ \begin{pmatrix} {0} &{} \quad {e} &{} \quad {0} \\ {e} &{} \quad {f} &{} \quad {0} \\ {0} &{} \quad {g} &{} \quad {0} \end{pmatrix},\ \begin{pmatrix} {0} &{} \quad {e} &{} \quad {0} \\ {e} &{} \quad {f} &{} \quad {g} \\ {0} &{} \quad {0} &{} \quad {0} \end{pmatrix} \ \text{ and }\ \begin{pmatrix} {0} &{} \quad {0} &{} \quad {0} \\ {0} &{} \quad {e} &{} \quad {0} \\ {0} &{} \quad k{e} &{} \quad {e} \end{pmatrix},} \end{aligned}$$

respectively, where H is a subset of K such that \(K = H\cup -H\) and \(H\cap -H = \{0\}\) (see [3] for details). We determine the (weak) multipliers of them below.

(0) \(A = C_0\) is the zero algebra, so by Proposition 2.5, we have

$$\begin{aligned} M'(A) = {A}^{A}, \ \,M(A) = \{T \in {A}^{A}\,|\, T(0) = 0 \}\ \text{ and }\ \, LM(A) = LM'(A) = L(A). \end{aligned}$$

(i) The unital algebras \(U_0, U_2\), \(U_3, U_4\) are commutative, so for such A we have

$$\begin{aligned} M(A) = LM(A) = M'(A) = LM'(A) = \{ l_x | x \in A \} \cong A \end{aligned}$$

by Theorem 5.3. For \(A = U_1\), we have

$$\begin{aligned} M(A) = LM(A) = M'(A) = LM'(A) \cong Z(A) = Ke. \end{aligned}$$

(ii) For \(A = C_1\), we observe that \(A_0 = \text{ Ann}_l(A) = \text{ Ann}_r(A) =Ke\), and we have a nihil decomposition \(A = A_1 \oplus A_0\) with \(A_1 = Kf + Kg\). Let \(T_1 \in M'_1(A)\), then by Theorem 3.1, \(T_1\) is a linear mapping such that \(T_1(Ke) = \{0\}\). Let

$$\begin{aligned} T_1 = \begin{pmatrix}0&{} \quad 0&{} \quad 0\\ 0&{} \quad q&{} \quad r\\ 0&{} \quad t&{} \quad u\end{pmatrix} \end{aligned}$$
(22)

with \(q,r,t,u \in K\) be the representation matrix of \(T_1\) on E. By Theorem 4.2, \(T_1\) is a weak multiplier if and only if

$$\begin{aligned} \begin{pmatrix} 0&{} \quad 0&{} \quad 0\\ 0&{} \quad te&{} \quad ue\\ 0&{} \quad -qe&{} \quad -re \end{pmatrix} = \varvec{A}T_1 = T_1^\textrm{t}\varvec{A} = \begin{pmatrix} 0&{} \quad 0&{} \quad 0\\ 0&{} \quad -te&{} \quad qe\\ 0&{} \quad -ue&{} \quad re \end{pmatrix}, \end{aligned}$$

if and only if \(r = t = 0\) and \(q = u\). Hence, \(M'_1(A) = \{T_q\,\big |\,q \in K\}\), where \(T_q = {\begin{pmatrix} 0&{} \quad 0&{} \quad 0\\ 0&{} \quad q&{} \quad 0\\ 0&{} \quad 0&{} \quad q \end{pmatrix}}\). By Theorem 3.1 we see

$$\begin{aligned} M'(A) = \{T_q\,\big |\,q \in K\} \oplus (Ke)^A \end{aligned}$$

and

$$\begin{aligned} LM'(A) = \left\{ {\begin{pmatrix} a&{} \quad b&{} \quad c\\ 0&{} \quad q&{} \quad 0\\ 0&{} \quad 0&{} \quad q \end{pmatrix}}\, \Big |\,a,b,c,q \in K\right\} . \end{aligned}$$

By examining the multiplication table of A, we find that  \(\alpha \beta = (xv-yz)e\)   for \(\alpha = xf+yg, \beta = zf+vg \in A_1\) with \(x,y,z,v \in K.\) By Corollary 3.2, \(T \in M'(A)\) is given by \(T = T_q + R\) with \(R \in (Ke)^A\) and this T is a multiplier if and only if

$$\begin{aligned} R((xv-yz)e)= & {} R(\alpha \beta )\ =\ \alpha T_q(\beta ) - T_q(\alpha \beta )\\ {}= & {} \alpha (q\beta ) - T_q((xv-yz)e)\ =\ q(xv-yz)e \end{aligned}$$

for any \(\alpha \) and \(\beta \), if and only if \(R(xe) = qxe\) for all \(x \in K\). Let \(S_q\) be the scalar multiplication in A by \(q \in K\). Then, we see

$$\begin{aligned} (T-S_q)(A) = (T_q - S_q + R)(A) \subseteq A_0 = Ke \end{aligned}$$

and

$$\begin{aligned} (T-S_q)(xe) = T_q(xe) + R(xe) - S_q(xe) = 0 + qxe -qxe = 0 \end{aligned}$$

for any \(x \in K\), that is, \((T - S_q)(Ke)= \{0\}\). Thus, we conclude

$$\begin{aligned} M(A) = \{S_q\,\big |\,q \in K\} \oplus \big \{R \in (Ke)^A\,\big |\,R(Ke) = \{0\}\big \}, \end{aligned}$$

and

$$\begin{aligned} LM(A) = \left\{ {\begin{pmatrix} a&{} \quad b&{} \quad c\\ 0&{} \quad a&{} \quad 0\\ 0&{} \quad 0&{} \quad a \end{pmatrix}}\, \Big |\,a,b,c \in K\right\} . \end{aligned}$$

(iii) \(A = C_2\): Because Ann\(_l(A) = Ke\) and Ann\(_r(A) = Kg\), we see \(A_0 = \{0\}\). Hence, any weak multiplier T is a linear multiplier by Proposition 5.1. By Theorem 4.2,

$$\begin{aligned} T =\begin{pmatrix} a&{} \quad b&{} \quad c\\ p&{} \quad q&{} \quad r\\ s&{} \quad t&{} \quad u \end{pmatrix} \end{aligned}$$
(23)

with \(a,b,c,p,q,r,s,t,u \in K\) is a (weak) multiplier if and only if

$$\begin{aligned} \begin{pmatrix} 0&{} \quad 0&{} \quad 0\\ ae+pf&{} \quad be+qf&{} \quad ce+rf\\ pg&{} \quad qg&{} \quad rg \end{pmatrix} = \varvec{A}T = T^\textrm{t}\varvec{A} = \begin{pmatrix} pe&{} \quad pf+sg&{} \quad 0\\ qe&{} \quad qf+tg&{} \quad 0\\ re&{} \quad rf+ug&{} \quad 0 \end{pmatrix}, \end{aligned}$$

if and only if \(b = c = p = r = s = t = 0\) and \(a = q = u\), that is, T is the scalar multiplication \(S_a\) by a. Consequently,

$$\begin{aligned} M(A) = M'(A) = LM(A) = LM'(A) = \{S_a\,\big |\,a \in K\} \cong K. \end{aligned}$$

(iv) \(C_3\) and \(C_4\) are opposite to each other, and share the same (weak) multipliers by Proposition 2.6. Let \(A = C_3\), then, A has a left identity g, \(Z_l(A) = A\) and Ann\(_l(A) = Ke+Kf\). Hence, by Theorem 5.3,

$$\begin{aligned} M(A) = M'(A) = LM(A) = LM'(A) = A/(Ke+Kg) = \{S_a\,\big |\,a \in K\}. \end{aligned}$$

(v) \(A = S_1\): We have \(A_0 = \text{ Ann}_l(A) = \text{ Ann}_r(A) = Kg\), and \(A = A_1 \oplus A_0\) with \(A_1 = Ke + Kf\). Then, \(T_1 \in M'_1(A)\) is a linear mapping with \(T(Kg) = \{0\}\). Let

$$\begin{aligned} T_1 = \begin{pmatrix} a&{} \quad b&{} \quad 0\\ p&{} \quad q&{} \quad 0\\ 0&{} \quad 0&{} \quad 0 \end{pmatrix} \end{aligned}$$
(24)

with \(a,b,p,q \in K\) be its representation on E. \(T_1\) is a weak multiplier if and only if

$$\begin{aligned} \begin{pmatrix} af+pg&{} \quad bf+qg&{} \quad 0\\ ag&{} \quad bg&{} \quad 0\\ 0&{} \quad 0&{} \quad 0 \end{pmatrix} = \varvec{A}T_1 = T_1^\textrm{t}\varvec{A} = \begin{pmatrix} af+pg&{} \quad ag&{} \quad 0\\ bf+qg&{} \quad bg&{} \quad 0\\ 0&{} \quad 0&{} \quad 0 \end{pmatrix}, \end{aligned}$$

if and only if \(b = 0\) and \(a = q\). Hence,

$$\begin{aligned} M'(A) = \left\{ T_1^{a,p}\,|\,a,p \in K\right\} \oplus (Kg)^A\ \, \text{ with } \ \,T_1^{a,p} = \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ p&{} \quad a&{} \quad 0\\ 0&{} \quad 0&{} \quad 0\end{pmatrix}}. \end{aligned}$$

So, \(T \in M'(A)\) is written as \(T = T_1^{a,p} + R\) with \(R \in (Kg)^A\), and this T is multiplier if and only if

$$\begin{aligned} R(xzf+(xv+yz)g)= & {} R(\alpha \beta )\ =\ \alpha T_1^{a,p}(\beta ) - T_1^{a,p}(\alpha \beta )\\= & {} (xe+yf)(aze+(pz+av)f) - axzf\\= & {} (pxz+a(xv+yz))g \end{aligned}$$

for any \(\alpha = xe+yf, \beta = ze+vf \in A_1\) with \(x,y,z,v \in K\), if and only if  \(R(xf+yg) = (px+ay)g\)  for all \(x, y \in K\). Let \(T^{a,p} = \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ p&{} \quad a&{} \quad 0\\ 0&{} \quad p&{} \quad a\end{pmatrix}}\), then \((T-T^{a,p})(A) \subseteq Kg\), and

$$\begin{aligned} (T-T^{a,p})(xf+yg)= & {} \left( T_1^{a,p}+R-T^{a,p}\right) (xf+yg)\\= & {} axf + (px+ay)g - (axf+pxg+ayg) = 0 \end{aligned}$$

for any \(x, y \in K\). Thus, \((T - T^{a,p})(Kf+Kg) = \{0\}\), and hence

$$\begin{aligned} M(A) = \{T^{a,p}\,|\,a, p \in K\} \oplus \big \{R \in (Kg)^A\,\big |\,R(Kf+Kg) = \{0\}\big \}. \end{aligned}$$

By intersecting \(M'(A)\) and M(A) with L(A), we obtain

$$\begin{aligned} LM'(A)= & {} \left\{ {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ p&{} \quad a&{} \quad 0\\ s&{} \quad t&{} \quad u\end{pmatrix}}\,\Big |\,a,p,s,t,u \in K\right\} \ \\{} & {} \text{ and }\ \, LM(A) = \left\{ {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ p&{} \quad a&{} \quad 0\\ s&{} \quad p&{} \quad a\end{pmatrix}}\,\Big |\,a,p,s \in K\right\} . \end{aligned}$$

(vi) \(A = S_2\): We have \(A_0 = \text{ Ann}_l(A) = \text{ Ann}_r(A) = Kg\), and \(A = A_1 \oplus A_0\) with \(A_1 = Ke + Kf\). Let a linear mapping \(T_1 \in M'_1(A)\) be represented as (24), then \(T_1\) is a weak multiplier if and only if

$$\begin{aligned} \begin{pmatrix}ae&{} \quad be&{} \quad 0\\ pg&{} \quad qg&{} \quad 0\\ 0&{} \quad 0&{} \quad 0\end{pmatrix} = \varvec{A}T = T^\textrm{t}\varvec{A} = \begin{pmatrix}ae&{} \quad pg&{} \quad 0\\ be&{} \quad qg&{} \quad 0\\ 0&{} \quad 0&{} \quad 0\end{pmatrix}, \end{aligned}$$

if and only if \(b = p = 0\). Hence,

$$\begin{aligned} M'(A) = \{T_1^{a,q}\,\big |\,a,q \in K\}\oplus (Kg)^A\ \,\text{ with }\ \, T_1^{a,q} = \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ 0&{} \quad q&{} \quad 0\\ 0&{} \quad 0&{} \quad 0\end{pmatrix}}, \end{aligned}$$

and

$$\begin{aligned} LM'(A) = \left\{ \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ 0&{} \quad q&{} \quad 0\\ s&{} \quad t&{} \quad u\end{pmatrix}}\,\Big |\,a,q,s,t,u \in K\right\} . \end{aligned}$$

By Corollary 3.2, a weak multiplier T written as \(T = T_1^{a,q} + R\) for \(a,q \in K\) and \(R \in (Kg)^A\) is multiplier if and only if

$$\begin{aligned} R(xze+yvg)= & {} R(\alpha \beta )\ =\ \alpha T_1^{a,q}(\beta ) -T_1^{a,q}(xze+yvg)\\= & {} (xe+yf)(aze+qvf) - axze \ =\ yqvg \end{aligned}$$

for any \(\alpha = xe+yf, \beta = ze+vf \in A_1\) with \(x,y,z,v \in K\), if and only if \(R(xe+yg) = qyg\) for all \(x, y \in K\). Let \(T^{a,q} = \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ 0&{} \quad q&{} \quad 0\\ 0&{} \quad 0&{} \quad q\end{pmatrix}}\), then we have \((T-T^{a,q})(xe+yg) = 0\) for any \(x, y \in K\), following the same calculation as in (v) above. Hence, \((T-T^{a,q})(Ke+Kg) = \{0\}\), and we have

$$\begin{aligned} M(A) = \left\{ T^{a,p}\,|\,a,p \in K\right\} \oplus \big \{R \in (Kg)^A\,\big |\,R(Ke+Kg) = \{0\}\big \} \end{aligned}$$

and

$$\begin{aligned} LM(A) = \left\{ \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ 0&{} \quad p&{} \quad 0\\ 0&{} \quad t&{} \quad 0\end{pmatrix}}\,\Big |\,a,p,t \in K\right\} . \end{aligned}$$

(vii) \(A = S_3\): We have \(A_0 = Kg\) and \(A = A_1 \oplus A_0\) with \(A_1 = Ke+Kf\). As \(A_1\) is a subalgebra of A, by Theorem 3.1 we obtain the equalities (10) in Sect. 3. Because \(A_1\) is a commutative unital algebra,

$$\begin{aligned} M(A_1) = M'(A_1) = A_1 = \left\{ \small {\begin{pmatrix}a&{} \quad 0\\ 0&{} \quad b\end{pmatrix}}\,\Big |\,a, b \in K\right\} \end{aligned}$$

by Theorem 5.3. Note that \(A^2 = Ke+Kf\). Hence,

$$\begin{aligned} M'(A) = A_1 \oplus (Kg)^A\ \,\text{ and }\ \, M(A) = A_1 \oplus \{T \in (Kg)^A\,|\,T(Ke+Kf) = 0\}. \end{aligned}$$

Intersecting with L(A) we have

$$\begin{aligned} LM'(A)= & {} \left\{ \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ 0&{} \quad b&{} \quad 0\\ s&{} \quad t&{} \quad u\end{pmatrix}}\, \Big |\,a,b,s,t,u \in K\right\} \ \\{} & {} \textrm{and}\ \ LM(A) = \left\{ \small {\begin{pmatrix}a&{} \quad 0&{} \quad 0\\ 0&{} \quad b&{} \quad 0\\ 0&{}0&{}u\end{pmatrix}}\,\Big |\,a,b,u \in K\right\} . \end{aligned}$$

(viii) \(A = S_4\): We have \(A = A_1 \oplus A_0\) with \(A_0 = Kg\) and \(A_1 = Ke+Kf\). Because \(A_1\) is a commutative unital subalgebra of A, similarly to the previous case, we obtain

$$\begin{aligned} M'(A) = A_1 \oplus (Kg)^A = \left\{ \small {\begin{pmatrix}a&{}0\\ b&{}a\end{pmatrix}}\,\Big |\,a, b \in K\right\} \oplus (Kg)^A, \end{aligned}$$
$$\begin{aligned}{} & {} M(A)=A_1 \oplus \big \{T \in (Kg)^A\,\big |\,T(Ke+Kf) = \{0\}\big \}, \\{} & {} LM'(A) \!=\! \left\{ \small {\begin{pmatrix}a&{}0&{}0\\ b&{}a&{}0\\ s&{}t&{}u\end{pmatrix}}\,\Big |\,a,b,s,t,u \!\in \! K\right\} \ \textrm{and}\ \ \!LM(A) \!=\! \left\{ \small {\begin{pmatrix}a&{}0&{}0\\ b&{}a&{}0\\ 0&{}0&{}u\end{pmatrix}}\,\Big |\,a,b,u \!\in \! K\right\} . \end{aligned}$$

(ix) \(A = W_1\): We have \(A = A_1 \oplus A_0\) with \(A_0 = Ke+Kf\) and \(A_1 = Kg\). Let \(T_1 \in M'_1(A)\), then \(T_1\) is a linear mapping with \(T_1(A_0) = \{0\}\). So \(T_1\) is determined by \(T_1(g) = ag\) with \(a \in K\). Denoting this \(T_1\) as \(T^a_1\), we have

$$\begin{aligned} M'(A) = \{T^a_1\,|\,a \in K\} \oplus (Ke+Kf)^A. \end{aligned}$$

A weak multiplier \(T = T^a_1 + R\) with \(R \in (Ke+Kf)^A\) is a multiplier if and only if

$$\begin{aligned} R(xye) = R((xg)(yg)) = xgT^a_1(yg) - T^a_1(xye) =axye \end{aligned}$$

for all \(x, y \in K\), if and only if \(R(xe) = axe\) for any \(x\in K\). Let \(T_a = \small {\begin{pmatrix}a&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad a\end{pmatrix}}\). Then, \((T-T_a)(Ke) = \{0\}\) and it follows that

$$\begin{aligned} M(A) = \{T_a\,\big |\,a\in K\} \oplus \big \{R \in (Ke+Kf)^A\,\big |\, R(Ke) = \{0\}\big \}. \end{aligned}$$

Also we have

$$\begin{aligned} LM'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ p&{}\quad q&{}\quad r\\ 0&{}\quad 0&{}\quad u\end{pmatrix}}\,\Big |\,a,b,c,p,q,r,u \in K\right\} \end{aligned}$$

and

$$\begin{aligned} LM(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ 0&{}\quad q&{}\quad r\\ 0&{}\quad 0&{}\quad a\end{pmatrix}}\,\Big |\,a,b,c,q,r \in K\right\} . \end{aligned}$$

(x) \(A = W_2\)Footnote 4: We have \(A = A_1 \oplus A_0\) with \(A_0 = Ke\) and \(A_1 = Kf+Kg\). Let \(T \in M'_1(A)\), then T is a linear mapping with \(T(Ke) = \{0\}\) and can be represented as (22). Then, T is a weak multiplier if and only if

$$\begin{aligned} \begin{pmatrix}0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad qe&{}\quad re\end{pmatrix} = \varvec{A}T = T^\textrm{t}\varvec{A} = \begin{pmatrix}0&{}\quad 0&{}\quad 0\\ 0&{}\quad te&{}\quad 0\\ 0&{}\quad ue&{}\quad 0\end{pmatrix}, \end{aligned}$$

if and only if \(r = t = 0\) and \(q = u\). Hence,

$$\begin{aligned} M'(A) = \{T_q\,|\,q \in K\}\oplus (Ke)^A\ \,\text{ with }\ \, T_q = \small {\begin{pmatrix}0&{}\quad 0&{}\quad 0\\ 0&{}\quad q&{}\quad 0\\ 0&{}\quad 0&{}\quad q\end{pmatrix}}. \end{aligned}$$

So,

$$\begin{aligned} LM'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ 0&{}\quad q&{}\quad 0\\ 0&{}\quad 0&{}\quad q\end{pmatrix}}\,\Big |\,a,b,c,q \in K\right\} . \end{aligned}$$

A weak multiplier \(T = T_q+R\) with \(R \in (Ke)^A\) is a multiplier if and only if

$$\begin{aligned} R(yze) = R(\alpha \beta ) = \alpha T_q(\beta ) - T_q(yze) = \alpha (q\beta ) = qyze \end{aligned}$$

for any \(\alpha = xf+yg, \beta = zf+vg \in A_1\) with \(x,y,z,v \in K\), if and only if \(R(xe) = qxe\) for all \(x \in K\). Let \(S_a\) be the scalar multiplication by \(a \in K\). Then, \((T-S_a)(Ke) = \{0\}\), and hence,

$$\begin{aligned} M(A) = \{S_a\,|\,a \in K\} \oplus \big \{R \in (Ke)^A\,\big |\,R(Ke) = \{0\}\big \} \end{aligned}$$

and

$$\begin{aligned} LM(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ 0&{}\quad a&{}\quad 0\\ 0&{}\quad 0&{}\quad a\end{pmatrix}}\,\Big |\,a,b,c \in K\right\} . \end{aligned}$$

(xi) \(A = W_4\): We have \(A = A_1 \oplus A_0\) with \(A_0 = Kf + Kg\) and \(A_1 = Ke\). Because \(A_1\) is a subalgebra isomorphic to the base field K, with the scalar multiplication \(S_1^a\) by \(a\in K\), we see

$$\begin{aligned} M'(A) = \{S_1^a\,|\,a \in K\} \oplus (fK+gK)^A \end{aligned}$$

and

$$\begin{aligned} M(A) = \{S_1^a\,|\,a \in K\} \oplus \big \{R \in (fK+gK)^A\,\big |\,R(Ke) = \{0\}\big \} \end{aligned}$$

by Theorem 3.1. Taking the intersections with L(A) we have

$$\begin{aligned} LM'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad 0&{}\quad 0\\ p&{}\quad q&{}\quad r\\ s&{}\quad t&{}\quad u\end{pmatrix}}\,\Big |\,a,p,q,r,s,t,u \in K\right\} \end{aligned}$$

and

$$\begin{aligned} LM(A) = \left\{ \small {\begin{pmatrix}a&{}\quad 0&{}\quad 0\\ 0&{}\quad q&{}\quad r\\ 0&{}\quad t&{}\quad u\end{pmatrix}}\,\Big |\,a,q,r,t,u \in K\right\} . \end{aligned}$$

(xii) \(W_5\) and \(W_6\) are opposite. Let \(A = W_5\), then \(A = A_1\oplus A_0\) with \(A_0 = Ke\) and \(A_1 = Kf+Kg\). Since \(A_1\) is a subalgebra of A and has a left identity g, we have

$$\begin{aligned} M(A_1) = LM(A_1) = M'(A_1) = LM'(A_1) \cong (A_1)/Kf \cong Kg \end{aligned}$$

by Theorem 5.3. So, any element in \(M(A_1)\) is a scalar multiplication \(S_1^q\) in \(A_1\) by \(q \in K\). By Theorem 3.1 we have

$$\begin{aligned}{} & {} M'(A) = \{S_1^q\,|\,q \in K\} \oplus (Ke)^A, \\{} & {} M(A) = \{S_1^q\,|\,q \in K\} \oplus \big \{R \in (Ke)^A\,\big |\,R(Kf+Kg) = \{0\}\big \}, \\{} & {} LM'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ 0&{}\quad q&{}\quad 0\\ 0&{}\quad 0&{}\quad q\end{pmatrix}}\,\Big |\,a,b,c,q \in K\right\} \ \, \textrm{and}\ \, LM(A) = \left\{ \small {\begin{pmatrix}a&{}\quad 0&{}\quad 0\\ 0&{}\quad q&{}\quad 0\\ 0&{}\quad 0&{}\quad q\end{pmatrix}}\,\Big |\,a,q \in K\right\} . \end{aligned}$$

(xiii) \(W_7\) and \(W_8\) are opposite. Let \(A = W_7\), then we see \(A_0 = \text{ Ann}_r(A) = \{0\}\). Hence, any weak multiplier is a linear multiplier by Proposition 5.1, and a linear mapping T represented as (23) is a weak multiplier if and only if

$$\begin{aligned} \begin{pmatrix}ae&{}\quad be&{}\quad ce\\ 0&{}\quad 0&{}\quad 0\\ pf+sg&{}\quad qf+tg&{}rf+ug\end{pmatrix} = \varvec{A}T = T^\textrm{t}\varvec{A} = \begin{pmatrix}ae&{}\quad sf&{}\quad sg\\ be&{}\quad tf&{}\quad tg\\ ce&{}\quad uf&{}\quad ug\end{pmatrix}, \end{aligned}$$

if and only if \(b = c = p = r = s = t = 0\) and \(q = u\). Therefore,

$$\begin{aligned} M(A) = LM(A) = M'(A) = LM'(A) = LM'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad 0&{}\quad 0\\ 0&{}\quad q&{}\quad 0\\ 0&{}\quad 0&{}\quad q\end{pmatrix}}\,\Big |\,a,q \in K\right\} . \end{aligned}$$

(xiv) \(W_9\) and \(W_{10}\) are opposite. Let \(A = W_9\). Then, because \(A_0 = \text{ Ann}_l(A) = \{0\}\), any weak multiplier is a linear multiplier and a linear mapping T represented as (23) is a weak multiplier if and only if

$$\begin{aligned} \begin{pmatrix}pe&{}\quad qe&{}\quad re\\ ae+pf&{}\quad be+qf&{}\quad ce+rf\\ pg&{}\quad qg&{}\quad rg\end{pmatrix} = \varvec{A}T = T^\textrm{t}\varvec{A} = \begin{pmatrix}pe&{}\quad ae+pf+sg&{}\quad 0\\ qe&{}\quad be+qf+tg&{}\quad 0\\ re&{}\quad ce+rf+ug&{}\quad 0\end{pmatrix}, \end{aligned}$$

if and only if \(c = p = r = s = t = 0, a = q = u\). Therefore,

$$\begin{aligned} LM(A) = M(A) = LM'(A) = M'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad 0\\ 0&{}\quad a&{}\quad 0\\ 0&{}\quad 0&{}\quad a\end{pmatrix}}\,\Big |\,a,b \in K\right\} . \end{aligned}$$

(xv) \(A = W_3(k)\): We have \(A = A_1 \oplus A_0\) with \(A_0 = Ke\) and \(A_1 = Kf+Kg\). Then, \(T \in M'_1(A)\) is a linear mapping with \(T(Ke) = \{0\}\) represented as (22). It is a weak multiplier if and only if

$$\begin{aligned} \begin{pmatrix}0&{}\quad 0&{}\quad 0\\ 0&{}\quad qe&{}\quad re\\ 0&{}\quad (kq+t)e&{}\quad (kr+u)e\end{pmatrix} = \varvec{A}T = T^\textrm{t}\varvec{A} = \begin{pmatrix}0&{}\quad 0&{}\quad 0\\ 0&{}\quad (q+kt)e&{}\quad te\\ 0&{}\quad (r+ku)e&{}\quad ue\end{pmatrix}. \end{aligned}$$
(25)

When \(k = 0\), (25) holds if and only if \(r = t\), and otherwise it holds if and only if \(r = t = 0\) and \(q =u\). Thus,

$$\begin{aligned} M'(A) = \{T_1^{q,r,u}\,\big |\,q,r,u \in K\} \oplus (Ke)^A,\ LM'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ 0&{}\quad q&{}r\\ 0&{}\quad r&{}\quad u\end{pmatrix}}\,\Big |\,a,b,c,q,r,u \in K\right\} \end{aligned}$$

when \(k = 0\), and

$$\begin{aligned} M'(A) = \{T_1^{q}\,\big |\,q \in K\} \oplus (Ke)^A,\ \, LM'(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ 0&{}\quad q&{}\quad 0\\ 0&{}\quad 0&{}\quad q\end{pmatrix}}\,\Big |\,a,b,c,q \in K\right\} \end{aligned}$$

when \(k \ne 0\), where

$$\begin{aligned} T_1^{q,r,u} = \small {\begin{pmatrix}0&{}\quad 0&{}\quad 0\\ 0&{}\quad q&{}\quad r\\ 0&{}\quad r&{}\quad u\end{pmatrix}}\ \, \textrm{and}\ \, T_1^{q} = \small {\begin{pmatrix}0&{}\quad 0&{}\quad 0\\ 0&{}\quad q&{}\quad 0\\ 0&{}\quad 0&{}\quad q\end{pmatrix}}. \end{aligned}$$

Now, when \(k = 0\), \(T = T_1^{q,r,u} + R\) with \(R \in (Ke)^A\) is multiplier if and only if

$$\begin{aligned} R((xz+yv)e)= & {} R(\alpha \beta )\ =\ \alpha T_1^{q,r,u}(\beta ) - T_1^{q,r,u}((xz+yv)e)\\= & {} \alpha ((qz+rv)f+(rz+uv)g) = (qxz+uyv+r(xv+yz))e \end{aligned}$$

for any \(\alpha = xf+yg, \beta = zf+vg \in A_1\) with \(x,y,z,v \in K\), if and only if \(q = u, r = 0\) and \(R(xe) = qxe\) for all \(x \in K\). While, when \(k \ne 0\), \(T = T_1^q + R\) with \(R \in (Ke)^A\) is a multiplier if and only if

$$\begin{aligned} R((xz + y(kz+v))e)= & {} R(\alpha \beta )\ =\ \alpha T_1^q(\beta ) - T_1^q(xue+y(kz+v)e)\\= & {} \alpha (q\beta )\ =\ q(xz + y(kz+v))e \end{aligned}$$

for any \(\alpha , \beta \), if and only if \(R(xe) = qxe\) for any \(x \in K\). In both cases, with the scalar multiplication \(S_a\) by \(a \in K\), we have  \((T-S_a)(Ke) = \{0\}\). Therefore, for arbitrary k (whether k is zero or nonzero) we conclude that

$$\begin{aligned} M(A) = \{S_a\,\big |\, a \in K\} \oplus \big \{R \in (Ke)^A\,\big |\,R(Ke) = \{0\}\big \} \end{aligned}$$

and

$$\begin{aligned} LM(A) = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad c\\ 0&{}\quad a&{}\quad 0\\ 0&{}\quad 0&{}\quad a\end{pmatrix}}\,\Big |\,a,b,c \in K\right\} . \end{aligned}$$

7 3-dimensional zeropotent algebras

An algebra A is zeropotent if \(x^2 = 0\) for all \(x \in A\). A zeropotent algebra A is anti-commutative, that is, \(xy = -yx\) for all \(x,y \in A\). Thus we see

$$\begin{aligned} A_0 = \text{ Ann}_l(A) = \text{ Ann}_r(A). \end{aligned}$$

Let A be a zeropotent algebras of dimension 3 over a field K with char\((K) \ne 2\). Let \({E} = \{e, f, g\}\) be a basis of A. Because A is anti-commutative, the multiplication table \(\varvec{A}\) of A on E is given as

$$\begin{aligned} \varvec{A} = \begin{pmatrix}0&{}\quad \alpha &{}\quad -\beta \\ alpha&{}\quad 0&{}\quad \gamma \\ \beta &{}\quad -\gamma &{}\quad 0\end{pmatrix} \ \ \text{ with }\ \ {\left\{ \begin{array}{ll} \gamma = fg = a_{11}e + a_{12}f + a_{13}g\\ \beta = ge = a_{21}e + a_{22}f + a_{23}g\\ \alpha = ef = a_{31}e + a_{32}f + a_{33}g \end{array}\right. } \end{aligned}$$

for \(a_{ij} \in K\). We call \(\small {\begin{pmatrix}a_{11}&{}\quad a_{12}&{}\quad a_{13}\\ a_{21}&{}\quad a_{22}&{}\quad a_{23}\\ a_{31}&{}\quad a_{32}&{}\quad a_{33}. \end{pmatrix}}\) the structural matrix of A. The rank of A is defined as the rank of its structural matrix.

Lemma 7.1

If \(\textrm{rank}(A) \ge 2\), then \(A_0 = \{0\}\).

Proof

If \(\textrm{rank}(A) \ge 2\), at least two of \(\alpha = ef, \beta = ge, \gamma = fg\) are linearly independent. Suppose that \(\alpha \) and \(\beta \) are linearly independent (the other cases are similar). If \(x = ae +bf +cg\) with \(a,b,c \in K\) is in \(\text{ Ann}_l(A)\), then \(xe = -b\alpha + c\beta \) and \(xf = a\alpha -c\gamma \) are both zero. It follows that \(a = b = c = 0\). Hence, we have \(A_0 = \text{ Ann}_l(A) = \{0\}\).

\(\square \)

Theorem 7.2

Let A be a zeropotent algebra of dimension 3 with rank\((A) \ge 2\) over K. Then, any weak multiplier of A is the scalar multiplication \(S_a\) for some \(a \in K\), that is,

$$\begin{aligned} M(A) = M'(A) = LM(A) = LM'(A) = \{S_a\,|\,a \in K\}. \end{aligned}$$

Proof

By Lemma 7.1 and Corollary 2.3, any weak multiplier T is a linear mapping. Let \(T \in L(A)\) be represented as (23). By Theorem 4.2, T is a weak multiplier if and only if \(\varvec{A}T = T^\textrm{t} \varvec{A}\), if and only if

$$\begin{aligned} {\begin{pmatrix}p\alpha -s\beta &{}\quad q\alpha -t\beta &{}\quad r\alpha -u\beta \\ -a\alpha +s\gamma &{}\quad -b\alpha +t\gamma &{}\quad -c\alpha +u\gamma \\ a\beta -p\gamma &{}\quad b\beta -q\gamma &{}\quad c\beta -r\gamma \end{pmatrix} = \begin{pmatrix}-p\alpha +s\beta &{}\quad a\alpha -s\gamma &{}\quad -a\beta +p\gamma \\ -q\alpha +t\beta &{}\quad b\alpha -t\gamma &{}\quad -b\beta +q\gamma \\ -r\alpha +u\beta &{}\quad c\alpha -u\gamma &{}\quad -c\beta +r\gamma \end{pmatrix}}\nonumber \\ \end{aligned}$$
(26)

holds. Suppose that \(\alpha , \beta \) are linearly independent (the other cases are similar). Then, by comparing the (1,1)-elements of the two matrices in (26), we have \(p\alpha - s\beta = -p\alpha +s\beta \), which implies \(p = s = 0\). Comparing the (1,2)-elements and (1,3)-elements, we have \(q\alpha - t\beta = a\alpha -s\gamma = a\alpha \) and \(r\alpha -u\beta = -a\beta +p\gamma = -a\beta \). It follows that \(a = q = u\) and \(r = t = 0\). Furthermore, comparing (2,2)-elements and (3,3)-elements, we see \(b = c = 0\). Consequently, (26) holds if and only if  \( b = c = p = r = s = t = 0\ \ \text{ and }\ \ a = q = u\), that is, \(T = S_a\). \(\square \)

In [4] we classified the zeropotent algebras of dimension 3 over an algebraically field K of characteristic not equal to 2. Up to isomorphism, we have 10 families of zeropotent algebras. They are

$$\begin{aligned} Z_0, Z_1, Z_2, Z_3, \{Z_4(a)\}_{a\in H}, Z_5, Z_6, \{Z_7(a)\}_{a\in H}, Z_8 \text{ and } Z_9 \end{aligned}$$

defined respectively by the structural matrices

$$\begin{aligned}{ \begin{pmatrix} 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0 \end{pmatrix}, \begin{pmatrix} 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix}, \begin{pmatrix} 0&{}\quad 0&{}\quad 1\\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix}, \begin{pmatrix} 0&{}\quad 1&{}\quad 0\\ -1&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 0 \end{pmatrix}, \begin{pmatrix} 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 1&{}\quad a\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix},} \end{aligned}$$
$$\begin{aligned} \begin{pmatrix} 0&{}\quad 1&{}\quad 0\\ 0&{}\quad 0&{}\quad 0\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix}, \begin{pmatrix} 0&{}\quad 1&{}\quad 1\\ 0&{}\quad 0&{}\quad 1\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix}, \begin{pmatrix} 1&{}\quad a&{}\quad 0\\ 0&{}\quad 1&{}\quad 0\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix}, \begin{pmatrix} 1&{}\quad 2&{}\quad 2\\ 0&{}\quad 1&{}\quad 2\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix} \text{ and } \begin{pmatrix} 1&{}\quad 3&{}\quad 3\\ 0&{}\quad 1&{}\quad 3\\ 0&{}\quad 0&{}\quad 1 \end{pmatrix}. \end{aligned}$$

\(Z_0\) is the zero algebra, and \(Z_1\) is isomorphic to the 3-dimensional associative algebra \(C_1\), and their (weak) multipliers are already determined in Sect. 6. The algebras \(Z_3\) to \(Z_9\) have rank greater or equal to 2, and they are covered by Theorem 7.2.

Thus, only \(A = Z_2\) remains to be analyzed. The multiplication table \(\varvec{A}\) of A is \({\begin{pmatrix} 0&{}\quad g&{}\quad 0\\ -g&{}\quad 0&{}\quad g\\ 0&{}\quad -g&{}\quad 0\end{pmatrix}}\). We see \(A_0 = \text{ Ann}_l(A) = \text{ Ann}_r(A) = K(e+g),\) and we have the nihil decomposition \(A = A_0 \oplus A_1\) with \(A_1 = Ke + Kf\). A weak multiplier \(T \in M'_1(A)\) is a linear mapping represented by \({\begin{pmatrix} a&{}\quad b&{}\quad c\\ p&{}\quad q&{}\quad r\\ 0&{}\quad 0&{}\quad 0 \end{pmatrix}}\) satisfying

$$\begin{aligned} \begin{pmatrix} pg&{}\quad qg&{}\quad rg\\ -ag&{}\quad -bg&{}\quad -cg\\ -pg&{}\quad -qg&{}\quad -rg \end{pmatrix} = \varvec{A}T = T^\textrm{t}\varvec{A} = \begin{pmatrix} -pg&{}\quad ag&{}\quad pg\\ -qg&{}\quad bg&{}\quad qg\\ -rg&{}\quad cg&{}\quad rg\end{pmatrix}. \end{aligned}$$

Hence, \(a = -c= q\) and \(b = p = r = 0\). Let \(T_a\) be this linear mapping, then by Theorem 3.1 we have

$$\begin{aligned} M'(A) = \{T_a\,|\,a \in K\} \oplus (K(f+g))^A \end{aligned}$$

and

$$\begin{aligned} LM'(A) = \left\{ {\begin{pmatrix} a+s&{}\quad t&{}\quad -a+u\\ 0&{}\quad a&{}\quad 0\\ s&{}\quad t&{}\quad u \end{pmatrix}}\, \Big |\,a,s,t,u \in K\right\} . \end{aligned}$$

By Corollary 3.2, a weak multiplier \(T = T_a + R\) with \(R \in (K(e+g))^A\) becomes a multiplier if and only if for any \(\zeta = xe+yf\) and \(\eta = ze+vf\) with \(x, y, z, v \in K\),

$$\begin{aligned} R((xv-yz)g)= & {} R(\zeta \eta )\ =\ \zeta T_a(\eta ) - T_a((xv-yz)g)\\= & {} \zeta (a\eta ) + a(xv-yz)e\ =\ a(xv-yz)(e+g) \end{aligned}$$

holds. It follows that \(R(xg) = ax(e+g)\) for all \(x \in K\). Let \(S_a\) be the scalar multiplication by a, then \((T-S_a)(A) \subseteq K(e+g)\) and \((T - S_a)(Kg) = \{0\}\). Hence, we obtain

$$\begin{aligned} M(A) = \{S_a\,|\,a \in K\} \oplus \big \{R \in (K(e+g))^A\,\big |\, R(Kg) = \{0\}\big \}, \end{aligned}$$

and

$$\begin{aligned} LM(A) = \left\{ \small {\begin{pmatrix}a+s&{}\quad t&{}\quad 0\\ 0&{}\quad a&{}\quad 0\\ s&{}\quad t&{}\quad a\end{pmatrix}}\,\Big |\,a,s,t \in K\right\} = \left\{ \small {\begin{pmatrix}a&{}\quad b&{}\quad 0\\ 0&{}\quad c&{}\quad 0\\ a-c&{}\quad b&{}\quad c\end{pmatrix}}\,\Big |\,a,b,c \in K\right\} . \end{aligned}$$